You are on page 1of 14

J. Astrophys. Astr.

(September 2017) 38:43 © Indian Academy of Sciences


DOI 10.1007/s12036-017-9470-9

Review

Superfluidity and Superconductivity in Neutron Stars

N. CHAMEL

Institute of Astronomy and Astrophysics, Université Libre de Bruxelles, CP 226, Boulevard du Triomphe,
1050 Brussels, Belgium.
E-mail: nchamel@ulb.ac.be

MS received 15 April 2017; accepted 14 June 2017; published online 12 September 2017

Abstract. Neutron stars, the compact stellar remnants of core-collapse supernova explosions, are unique
cosmic laboratories for exploring novel phases of matter under extreme conditions. In particular, the occurrence
of superfluidity and superconductivity in neutron stars will be briefly reviewed.

Keywords. Neutron stars—superfluidity—superconductivity—dense matter.

1. Introduction With typical temperatures of order 107 K, the highly


degenerate matter in neutron stars is expected to become
Formed in the furnace of gravitational core-collapse cold enough for the appearance of superfluids and super-
supernova explosions of stars with a mass between 8 and conductors – frictionless quantum liquids respectively
10 times that of the Sun (Deshpande et al. 1995), neutron electrically neutral and charged (Leggett 2006) – made
stars contain matter crushed at densities exceeding that of neutrons and protons, and more speculatively of other
found inside the heaviest atomic nuclei (for a general particles such as hyperons or quarks. If these phase tran-
review about neutron stars, see, Srinivasan (1997) and sitions really occur, neutron stars would not only be the
Haensel et al. (2007)). A proto neutron star is initially largest superfluid and superconducting systems known
fully fluid with a mass of about one or two solar masses, in the Universe (Srinivasan 1997; Chamel & Haensel
a radius of about 50 km and internal temperatures of 2008; Sauls 1989; Sedrakian & Clark 2006; Page et al.
the order of 1011 –1012 K (for a review about neutron- 2014; Graber et al. 2017), but also the hottest ones with
star formation, see Prakash et al. (2001)). About one critical temperatures of the order of 1010 K as compared
minute later, the proto-neutron star becomes transparent to a mere 203 K for the world record achieved in 2014
to neutrinos that are copiously produced in its interior, in terrestrial laboratories and consisting of hydrogen-
thus rapidly cools down and shrinks into an ordinary sulphide compound under high pressure (Drozdov et al.
neutron star. After a few months, the surface of the 2015).
star – possibly surrounded by a very thin atmospheric After describing the main properties of terrestrial
plasma layer of light elements – still remains liquid. superfluids and superconductors, an overview of the the-
However, the layers beneath crystallize thus forming oretical developments in the modelling of superfluid and
a solid crust (Chamel & Haensel 2008). At this point, superconducting neutron stars will be given. Finally, the
the core is much colder than the crust because of the possible observational manifestations of these phases
cooling power of the escaping neutrinos. After several will be briefly discussed.
decades, the interior of the star reaches a thermal equi-
librium with temperatures of about 108 K (except for
a thin outer heat-blanketing envelope). The last cool- 2. Terrestrial superfluids and superconductors
ing stage takes place after about a hundred thousand
years, when heat from the interior diffuses to the sur- 2.1 Historical milestones
face and is dissipated in the form of radiation (for a
recent review about neutron-star cooling, see Potekhin Superconductivity and superfluidity were known long
et al. (2015)). before the discovery of pulsars in August 1967. Heike
43 Page 2 of 14 J. Astrophys. Astr. (September 2017) 38:43

Table 1. Properties of various superfluid and superconducting sys-


tems in order of their critical temperature Tc . Adapted from Table 1.1
in Leggett (2006).

System Density (cm−3 ) Tc (K)

Neutron stars ∼1039 ∼1010


Cuprates and other exotics ∼1021 1–165
Electrons in ordinary metals ∼1023 1–25
Helium-4 ∼1022 2.17
Helium-3 ∼1022 2.491 × 10−3
Fermi alkali gases ∼1012 ∼10−6
Bose alkali gases ∼1015 ∼10−7 –10−5

Kamerlingh Onnes and his collaborators were the first see Balibar (2007, 2014)). In particular, helium II does
to liquefy helium in 1908, thus allowing them to explore not boil, as was actually first noticed by Kamerlingh
the properties of materials at lower temperatures than Onnes and his collaborators the same day they dis-
could be reached before. On 8th April 1911, they covered superconductivity (van Delft & Kes 2010).
observed that the electric resistance of mercury dropped Helium II can flow without resistance through very nar-
to almost zero at temperature Tc  4.2 K (for an his- row slits and capillaries, almost independently of the
torical account of this discovery, see e.g., van Delft & pressure drop. The term ‘superfluid’ was coined by
Kes (2010)). Two years later, lead and tin were found Pyotr Kapitsa in 1938 by analogy with superconduc-
to be also superconducting. In 1914, Onnes showed tors (Kapitsa 1938). Helium II also flows up over the
that superconductivity is destroyed if the magnetic field sides of a beaker and drip off the bottom (for ordinary
exceeds some critical value. He later designed an exper- liquids, the so-called Rollin film is clamped by viscos-
iment to measure the decay time of a magnetically ity). The existence of persistent currents in helium II
induced electric current in a superconducting lead ring, was experimentally established at the end of the 1950s
and did not notice any change after an hour. Super- and the beginning of 1960s (Reppy & Depatie 1964).
conducting currents can actually be sustained for more The analog of the Meissner–Ochsenfeld phenomenon,
than hundred thousand years (File & Mills 1963). Kee- which was predicted by Fritz London, was first observed
som & Kok (1932) found that the heat capacity of by George Hess and William Fairbank at Stanford in
tin exhibits a discontinuity as it becomes supercon- June 1967 (Hess & Fairbank 1967): the angular momen-
ducting thus demonstrating that this phase transition tum of helium-4 in a slowly rotating container was found
is of second order. A year later, Meissner & Ochsen- to be reduced as the liquid was cooled below the critical
feld (1933) made the remarkable observation that when temperature Tλ .
a superconducting material initially placed in a mag- At the time the first observed pulsars were identi-
netic field is cooled below the critical temperature, fied as neutron stars, several materials had thus been
the magnetic flux is expelled from the sample. This found to be superconducting, while helium-4 was the
showed that superconductivity represents a new ther- unique superfluid known. The superfluidity of helium-
modynamical equilibrium state of matter. Rjabinin & 3 was established by Osheroff et al. (1972). No other
Shubnikov (1935a, b) at the Kharkov Institute of Sci- superfluids were discovered during the next two decades
ence and Technology in Ukraine discovered that some until the production of ultracold dilute gases of bosonic
so-called ‘hard’ or type II superconductors (as opposed atoms in 1995 (Anderson et al. 1995; Davis et al. 1995),
to ‘soft’ or type I superconductors) exhibit two critical and of fermionic atoms in 2003 (Regal et al. 2004). The
fields, between which the magnetic flux partially pen- main properties of some known superfluids and super-
etrates the material. Various superconducting materials conductors are summarized in Table 1.
were discovered in the following decades.
During the 1930s, several research groups in Lei-
den, Toronto, Moscow, Oxford and Cambridge (United 2.2 Quantum liquids
Kingdom), found that below Tλ  2.17 K, helium-
4 (referred to as helium II) does not behave like an Superconductivity and superfluidity are among the most
ordinary liquid (for a review of the historical context, spectacular macroscopic manifestations of quantum
J. Astrophys. Astr. (September 2017) 38:43 Page 3 of 14 43

mechanics. Satyendra Nath Bose and Albert Einstein where p is the momentum per superfluid particle, h
predicted in 1924–1925 that at low enough tempera- denotes the Planck’s constant, N is any integer, and
tures an ideal gas of bosons condense into a macroscopic the integral is taken over any closed path. It can be
quantum state (Bose 1924; Einstein 1925). The associ- immediately recognized that this condition is the Bohr–
ation between Bose–Einstein condensation (BEC) and Sommerfeld quantisation rule. The flow quantisation
superfluidity was first advanced by London (1938). The follows from the fact that a superfluid is a macroscopic
only known superfluid at the time was helium-4, which quantum system whose momentum is thus given by
is a boson. The condensate can behave coherently on p = h/λ, where λ is the de Broglie wavelength. Requir-
a very large scale and can thus flow without any resis- ing the length of any closed path to be an integral
tance. It was a key idea for developing the microscopic multiple of the de Broglie wavelength leads to equa-
theory of superfluidity and superconductivity. Soon tion (1). The physical origin of this condition has been
afterwards, Tisza (1938) postulated that a superfluid usually obscured by the introduction of the ‘superfluid
such as He II contains two distinct dynamical compo- velocity’ V s = p/m, where m is the mass of the super-
nents: the condensate, which carries no entropy, coexists fluid particles.
with a normal viscous fluid. This model explained all In a rotating superfluid, the flow quantisation condi-
phenomena observed at the time and predicted thermo- tion (1) leads to the appearence of N quantised vortices.
mechanical effects like ‘temperature waves’. Although In a region free of vortices, the superflow is character-
Landau (1941) incorrectly believed that superfluidity is ized by the irrotationality condition
not related to BEC, he developed the two-fluid model
∇ × p = 0. (2)
and showed, in particular, that the normal fluid consists
of ‘quasiparticles’, which are not real particles but com- Inside a vortex, the superfluidity is destroyed. Because
plex many-body motions. This two-fluid picture was superfluid vortices are essentially of quantum nature,
later adapted to superconductors (Gorter 1955). their internal structure cannot be described by a purely
According to the microscopic theory of supercon- hydrodynamic approach. However, vortices can be
ductivity by John Bardeen, Leon Cooper and Robert approximately treated as structureless topological
Schrieffer (BCS) published in 1957 (Bardeen et al. defects at length scales much larger than the vortex core
1957), the dynamical distortions of the crystal lattice size. As shown by Tkachenko (1966), quantised vortices
(phonons) in a solid can induce an attractive effec- tend to arrange themselves on a regular triangular array,
tive interaction between electrons of opposite spins. with a spacing given by
Roughly speaking, electrons can thus form pairs and 
undergo a BEC below some critical temperature. A h
dυ = √ , (3)
superconductor can thus be viewed as a charged super- 3m
fluid. This picture however should not be taken too far. where  is the angular frequency. Vortex arrays have
Indeed, electron pairs are very loosely bound and over- been observed in superfluid helium (Yarmchuk et al.
lap. Their size ξ ∼ h̄vF /(kB Tc ) (usually referred to as 1979) and more recently in atomic Bose–Einstein con-
the coherence length), with vF the Fermi velocity, kB densates (Abo-Shaeer et al. 2001; Zwierlein et al.
Boltzmann’s constant, and Tc the critical temperature, 2005). At length scales much larger than the intervor-
is typically much larger than the lattice spacing. More- tex spacing dυ , the superfluid flow mimics rigid body
over, electron pairs disappear at temperatures T > Tc . rotation such that
The BEC and the BCS transition are now understood as
two different limits of the same phenomenon. The pair- ∇ × p = mn υ κ , (4)
ing mechanism suggested that fermionic atoms could where n υ is the surface density of vortices given by
also become superfluid, as was later confirmed by the
m
discovery of superfluid helium-3. Since 2003, various nυ = , (5)
other fermionic superfluids have been found, as men- π h̄
tioned in the previous section. and the vector κ, whose norm is equal to h/m, is aligned
As first discussed by Onsager (1949) and Feynman with the average angular velocity. Landau’s original
(1955), the quantum nature of a superfluid is embedded two-fluid model was further improved in the 1960s by
in the quantisation of the flow Hall & Vinen (1956), Hall (1960), and independently
by Bekarevich & Khalatnikov (1961) to account for the

presence of quantised vortices within a coarse-grained
p · d = N h , (1) average hydrodynamic description.
43 Page 4 of 14 J. Astrophys. Astr. (September 2017) 38:43

The quantisation condition (1) also applies to super- 3. Superstars


conductors. But in this case, the momentum (in CGS
units) is given by p ≡ m v + (q/c)A A where m, 3.1 Prelude: internal constitution of a neutron star
q, and v are the mass, electric charge and velocity
of superconducting particles respectively, and A is A few meters below the surface of a neutron star, matter
the electromagnetic potential vector. Introducing the is so compressed by the tremendous gravitational pres-
density n of superconducting particles and the ‘super- sure that atomic nuclei, which are supposedly arranged
current’ J = nqv, the situation N = 0 as described by on a regular crystal lattice, are fully ionised and thus
equation (2) leads to London’s equation coexist with a quantum gas of electrons. With increasing
c depth, nuclei become progressively more neutron-rich.
∇ ×J = − B, (6) Only in the first few hundred metres below the sur-
4π λ2L
face can the composition be completely determined by
where B = ∇ × A is the magnetic field induction, and experimentally measured masses of atomic nuclei (Wolf
λL = mc2 /(4π nq 2 ) is the London penetration depth. et al. 2013). In the deeper layers recourse must be made
Situations with N > 0 are encountered in type II super- to theoretical models (Pearson et al. 2011; Kreim et al.
conductors for which λL  ξ . Considering a closed 2013; Chamel et al. 2015; Sharma et al. 2015, Utama
contour outside a sample of such a superconductor for et al. (2016), Chamel et al. 2017). At densities of a
which J = 0 and integrating the momentum p along few 1011 g cm−3 , neutrons start to ‘drip’ out of nuclei
this contour, leads to the quantisation of the total mag- (see Chamel et al. (2015) for a recent discussion). This
netic flux into fluxoids (also referred to as flux tubes marks the transition to the inner crust, an inhomoge-
or fluxons) neous assembly of neutron-proton clusters immersed in
 an ocean of unbound neutrons and highly degenerate
= A · d = N 0 , (7) electrons. According to various calculations, the crust
dissolves into a uniform mixture of neutrons, protons
where 0 = hc/|q| is the flux quantum. The mag- and electrons when the density reaches about half the
netic flux quantization, first envisioned by London, was density ∼2.7 × 1014 g cm−3 found inside heavy atomic
experimentally confirmed in 1961 by Bascom Deaver nuclei (see Chamel & Haensel (2008) for a review
and William Fairbank at Stanford University (Deaver about neutron-star crusts). Near the crust-core interface,
& Fairbank 1961), and independently Robert Doll and nuclear clusters with very unusual shapes such as elon-
Martin Näbauer at the Low Temperature Institute in gated rods or slabs may exist (see section 3.3 of Chamel
Hersching (Doll & Näbauer 1961). As predicted by & Haensel (2008), see also Watanabe & Maruyama
Abrikosov (1957), these fluxoids tend to arrange them- (2012)). These so-called ‘nuclear pastas’ could account
selves into a triangular lattice with a spacing given by for half of the crustal mass, and play a crucial role for
the dynamical evolution of the star and its cooling (Pons
 et al. 2013; Horowitz et al. 2015). The composition of
2hc the innermost part of neutron-star cores remains highly
d = √ . (8)
3|q|B uncertain: apart from nucleons and leptons, it may also
contain hyperons, meson condensates, and deconfined
Averaging at length scales much larger than dυ , the sur- quarks (Haensel et al. (2007); see also Sedrakian (2010),
face density of fluxoids is given by Chatterjee & Vidaña (2016)).
B |q|B
n = = , (9) 3.2 Superfluid and superconducting phase transitions
0 hc
in dense matter
where B denotes the average magnetic field strength.
The size of a fluxoid (within which the superconductiv- Only one year after the publication of the BCS theory
ity is destroyed) is of the order of the coherence length of superconductivity, Bogoliubov (1958) was the first
ξ . The magnetic field carried by a fluxoid extends over to consider the possibility of superfluid nuclear matter.
a larger distance of the order of the London penetration Migdal (1959) speculated that the interior of a neutron
length λL . The nucleation of a single fluxoid thus occurs star might contain a neutron superfluid, and its critical
at a critical field Hc1 ∼ 0 /(π λ2L ), and superconduc- temperature was estimated by Ginzburg & Kirzhnits
tivity is destroyed at the critical field Hc2 ∼ 0 /(π ξ 2 ) (1964) using the BCS theory. Proton superconductiv-
at which point the cores of the fluxoids touch. ity in neutron stars was studied by Wolf (1966). The
J. Astrophys. Astr. (September 2017) 38:43 Page 5 of 14 43

possibility of anisotropic neutron superfluidity was 2006; Maurizio et al. 2014; Ding et al. 2016). At neutron
explored by Hoffberg et al. (1970), and independently concentrations above ∼0.16 fm−3 , pairing in the cou-
by Tamagaki (1970). pled 3 PF2 channel becomes favored but the maximum
Neutrons and protons are fermions, and due to the critical temperature remains very uncertain, predictions
Pauli exclusion principle, they generally tend to avoid ranging from ∼108 K to ∼109 K (Maurizio et al. 2014;
themselves. This individualistic behaviour, together Ding et al. 2016; Baldo et al. 1998; Dong et al. 2013).
with the strong repulsive nucleon–nucleon interaction at This lack of knowledge of neutron superfluid properties
short distance, provide the necessary pressure to coun- mainly stems from the highly nonlinear character of the
terbalance the huge gravitational pull in a neutron star, pairing phenomenon, as well as from the fact that the
thereby preventing it from collapsing. However at low nuclear interactions are not known from first principles
enough temperatures, nucleons may form pairs (Broglia (see Machleidt (2017) for a recent review).
& Zelevinsky 2013) similarly as electrons in ordi- Another complication arises from the fact that neu-
nary superconductors as described by the BCS theory1 . tron stars are not only made of neutrons. The presence of
These bosonic pairs can therefore condense, analo- nuclear clusters in the crust of a neutron star may change
gous to superfluid helium-3. While helium-3 becomes substantially the neutron superfluid properties. Unfor-
a superfluid only below 1 mK, nuclear superfluidity tunately, microscopic calculations of inhomogeneous
could be sustainable even at a temperature of several crustal matter employing realistic nuclear interactions
billions degrees in a neutron star due to the enormous are not feasible. State-of-the-art calculations are based
pressure involved. The nuclear pairing phenomenon is on the nuclear energy density functional theory, which
also supported by the properties of atomic nuclei (Dean allows for a consistent and unified description of atomic
& Hjorth-Jensen 2003). nuclei, infinite homogeneous nuclear matter and neu-
Because the nuclear interactions are spin-dependent tron stars (see Chamel et al. (2013) and references
and include non-central tensor components (angular therein). The main limitation of this approach is that
momentum-dependent), different kinds of nucleon– the exact form of the energy density functional is not
nucleon pairs could form at low enough temperatures. known. In practice, phenomenological functionals fit-
The most attractive pairing channels2 are 1 S0 at low ted to selected nuclear data must therefore be employed.
densities and the coupled 3 PF2 channel at higher densi- The superfluid in neutron-star crusts, which bears simi-
ties (Gezerlis et al. 2014). In principle, different types larities with terrestrial multiband superconductors, was
of pairs may coexist. However, one or the other are first studied within the band theory of solids in Chamel
usually found to be energetically favored (Lombardo et al. (2010). However, this approach is computationally
& Schulze 2001). Let us mention that nucleons may very expensive, and has been so far limited to the deep-
also form quartets such as α-particles, which can them- est layers of the crust. For this reason, most calculations
selves condense at low enough temperatures (Schuck of neutron superfluidity in neutron-star crusts (Mar-
2014). Most microscopic calculations have been car- gueron & Sandulescu 2012) have been performed using
ried out in pure neutron matter using diagrammatic, an approximation introduced by Wigner & Seitz (1933)
variational, and more recently, Monte Carlo methods in the context of electrons in metals: the Wigner–Seitz or
(see Gezerlis et al. 2014; Lombardo & Schulze 2001 Voronoi cell of the lattice (a truncated octahedron in case
for a review). At concentrations below ∼0.16 fm−3 , as of a body-centred cubic lattice) is replaced by a sphere
encountered in the inner crust and in the outer core of of equal volume. However, this approximation can only
a neutron star, neutrons are expected to become super- be reliably applied in the shallowest region of the crust
fluid by forming 1 S0 pairs, with critical temperatures of due to the appearance of spurious shell effects (Chamel
about 1010 K at most (Gezerlis et al. 2014; Cao et al. et al. 2007). Such calculations have shown that the phase
diagram of the neutron superfluid in the crust is more
1 The high temperatures ∼107 K prevailing in neutron star interiors complicated than that in pure neutron matter; in particu-
prevent the formation of electron pairs recalling that the highest lar, the formation of neutron pairs can be enhanced with
critical temperatures of terrestrial superconductors do not exceed increasing temperature (Margueron & Khan 2012; Pas-
∼200 K. In particular, iron expected to be present in the outermost tore et al. 2013; Pastore 2015). Microscopic calculations
layers of a neutron star was found to be superconducting in 2001,
in pure neutron matter at densities above the crust-core
but with a critical temperature Tc  2 K (Shimizu et al. 2001). See
also Ginzburg (1969). boundary are not directly applicable to neutron stars due
2 A given channel is denoted by 2S+1 L , where J is the total angular to the presence of protons, leptons, and possibly other
J
momentum, L is the orbital angular momentum, and S the spin of particles in neutron-star cores. Few microscopic calcu-
nucleon pair. lations have been performed so far in beta-stable matter
43 Page 6 of 14 J. Astrophys. Astr. (September 2017) 38:43

(Zhou et al. 2004). Because the proton concentration indicate that ordinary pulsars can also be endowed with
in the outer core of a neutron star is very low, pro- very high magnetic fields of order 1014 G (Ng & Kaspi
tons are expected to become superconducting in the 1 S0 2011). According to numerical simulations, neutron
channel. However, the corresponding critical tempera- stars may potentially be endowed with internal magnetic
tures are very poorly known due to the strong influence fields as high as 1018 G (see Pili et al. 2014; Chatterjee
of the surrounding neutrons (Baldo & Schulze 2007). et al. 2015 and references therein).
Neutron–proton pairing could also in principle occur, The presence of a high magnetic field in the interior of
but is usually disfavored by the very low proton con- a neutron star may have a large impact on the superfluid
tent of neutron stars (Stein et al. 2014). Other more and superconducting phase transitions. Proton super-
speculative possibilities include hyperon–hyperon and conductivity is predicted to disappear at a critical field
hyperon-nucleon pairing (Chatterjee & Vidaña (2016) of order 1016 –1017 G (Baym et al. 1969a). Because
and references therein). The core of a neutron star might spins tend to be aligned in a magnetic field, the forma-
also contain quarks in various color superconducting tion of neutron pairs in the 1 S0 channel is disfavored in
phases (Alford et al. 2008). a highly magnetised environment, as briefly mentioned
According to cooling simulations, the temperature in by Kirzhnits (1970). It has been recently shown that
a neutron star is predicted to drop below the estimated 1 S pairing in pure neutron matter is destroyed if the
0
critical temperatures of nuclear superfluid phases after magnetic field strength exceeds ∼1017 G (Stein et al.
∼10–102 years. The interior of a neutron star is thus 2016). Moreover, the magnetic field may also shift the
thought to contain at least three different kinds of super- onset of the neutron-drip transition in dense matter to
fluids and superconductors (Page et al. 2014): (i) an higher or lower densities due to Landau quantisation
isotropic neutron superfluid (with 1 S0 pairing) perme- of electron motion, thus changing the spatial extent of
ating the inner region of the crust and the outer core, (ii) the superfluid region in magnetar crusts (Chamel et al.
an anisotropic neutron superfluid (with 3 PF2 pairing) in 2015; Fantina et al. 2016; Basilico et al. 2015; Chamel
the outer core, and (iii) an isotropic proton supercon- et al. 2016).
ductor (with 1 S0 pairing) in the outer core. The neutron
superfluids in the crust and in the outer core are not
expected to be separated by a normal region.
3.4 Dynamics of superfluid and superconducting
neutron stars
3.3 Role of a high magnetic field
The minimal model of superfluid neutron stars consists
Most neutron stars that have been discovered so far are of at least two distinct interpenetrating dynamical com-
radio pulsars with typical surface magnetic fields of ponents (Baym et al. 1969): (i) a plasma of electrically
order 1012 G (as compared to ∼10−1 G for the Earth’s charged particles (electrons, nuclei in the crust and pro-
magnetic field), but various other kinds of neutron stars tons in the core) that are essentially locked together by
have been revealed with the development of the X-ray the interior magnetic field, and (ii) a neutron superfluid.
and gamma-ray astronomy (Harding 2013). In partic- Whether protons in the core are superconducting or not,
ular, a small class of very highly magnetised neutron they co-move with the other electrically charged parti-
stars thus dubbed magnetars by Thompson & Duncan cles (Sauls 1989).
(1992) (see Woods & Thompson (2006) for a review) The traditional heuristic approach to superfluid hydro-
have been identified in the form of soft-gamma ray dynamics blurring the distinction between velocity and
repeaters (SGRs) and anomalous X-ray pulsars (AXPs). momentum makes it difficult to adapt and extend Lan-
Tremendous magnetic fields up to about 2×1015 G have dau’s original two-fluid model to the relativistic context,
been measured at the surface of these stars from both as required for a realistic description of neutron stars
spin-down and spectroscopic studies (Olausen & Kaspi (Carter & Khalatnikov 1994). In particular, in super-
2014; Tiengo et al. 2013; An et al. 2014), and vari- fluid mixtures such as helium-3 and helium-4 (Andreev
ous observations suggest the existence of even higher & Bashkin 1976), or neutrons and protons in the core of
internal fields (Stella et al. 2005; Kaminker et al. 2007; neutron stars (Sedrakyan & Shakhabasyan 1980; Var-
Watts & Strohmayer 2007; Samuelsson & Andersson danyan & Sedrakyan 1981), the different superfluids
2007; Vietri et al. 2007; Rea et al. 2010; Makishima are generally mutually coupled by entrainment effects
et al. 2014). Although only 23 such stars are currently whereby the true velocity v X and the momentum pX of
known (Olausen & Kaspi 2014), recent observations a fluid X are not aligned:
J. Astrophys. Astr. (September 2017) 38:43 Page 7 of 14 43

pX = KXYvY , (10) lead to the discovery of new conservation laws in super-
Y fluid systems such as the conservation of generalised
helicy currents.
where KXY is a symmetric matrix determined by the As pointed out by Ginzburg & Kirzhnits (1964), the
interactions between the constituent particles. In the interior of a rotating neutron star is expected to be
two-fluid model, entrainment can be equivalently for- threaded by a very large number of neutron superfluid
mulated in terms of ‘effective masses’. Considering the vortices (for a discussion of the vortex structure in 1 S0
neutron–proton mixture in the core of neutron stars, and 3 PF2 neutron superfluids, see Sauls 1989). Intro-
the neutron momentum can thus be expressed as pn = ducing the spin period P in units of 10 ms, P10 ≡
m n vn in the proton rest frame (vv p = 0), with m n = Knn . P/(10 ms), the surface density of vortices (5) is of the
Alternatively, a different kind of effective mass can order
be introduced, namely m n = Knn − Knp K pn /K pp ,
−1
such that pn = m n vn in the proton momentum rest n υ ∼ 6 × 105 P10 cm−2 . (11)
frame (pp p = 0). These effective masses should not be The average intervortex spacing (3) of order
confused with those introduced in microscopic many- 
body theories (Chamel & Haensel 2006). Because of dυ ∼ n −1/2
υ ∼ 10 −3
P10 cm , (12)
the strong interactions between neutrons and protons,
is much larger than the size of the vortex core (Yu &
entrainment effects in neutron-star cores cannot be
Bulgac 2003). Neutron superfluid vortices can pin to
ignored (see Chamel & Haensel 2006; Gusakov &
nuclear inhomogeneities in the crust. However, the pin-
Haensel 2005; Chamel 2008; Kheto & Bandyopadhyay
ning strength remains uncertain (see Wlazłowski et al.
2014; Sourie et al. 2016 for recent estimates). In the
2016 and references therein; see also section 8.3.5 of
neutron-rich core of neutron stars, we typically have
Chamel & Haensel (2008)). Protons in the core of a
m n ∼ m n ∼ m n , and m p ∼ m p ∼ (0.5−1)m p , neutron star are expected to become superconducting at
where m n and m p denote the ‘bare’ neutron and pro- low enough temperatures. Contrary to superfluid neu-
ton masses respectively. As shown by Carter (1975), at trons, superconducting protons do not form vortices. As
the global scale of the star, general relativity induces shown by Baym et al. (1969a, b), the expulsion of the
additional couplings between the fluids due to Lense– magnetic flux accompanying the transition takes place
Thirring effects, which tend to counteract entrainment. on a very long time scale ∼106 years due to the very
As recently found in Sourie et al. (2017), frame- high electrical conductivity of the dense stellar matter.
dragging effects can be as important as entrainment. The superconducting transition thus occurs at constant
An elegant variational formalism to derive the hydro- magnetic flux. The proton superconductor is usually
dynamic equations of any relativistic (super)fluid mix- thought to be of type II (Baym et al. 1969a) (see also
tures was developed by Carter and collaborators (Carter Charbonneau & Zhitnitsky 2007; Alford et al. 2008 and
1989, 2001; Gourgoulhon 2006; Andersson & Comer references therein), in which case, the magnetic flux
2007). This formalism relies on an action principle penetrates the neutron star core by forming fluxoids,
in which the basic variables are the number densi- with a surface density (9) of order
ties and currents of the different fluids. The equations
of motion can be derived by considering variations n ∼ 5 × 1018 B12 cm−2 , (13)
of the fluid particle trajectories. Dissipative processes
where the magnetic field strength B is expressed as
(e.g. viscosity in non-superfluid constituents, super-
B12 ≡ B/(1012 G). This surface density corresponds
fluid vortex drag, ordinary resistivity between non-
to a spacing (8) of order
superfluid constituents, nuclear reactions) can be treated 
within the same framework. The convective formalism −1/2 −1
d ∼n ∼ 5 × 10−10 B12 cm . (14)
developed by Carter was later adapted to the com-
paratively more intricate Newtonian theory within a Since the magnetic flux is frozen in the stellar core, flux-
4-dimensionally covariant framework (see Carter & oids can form even if the magnetic field is lower than the
Chamel 2004, 2005a, b; see also Prix 2004, 2005; critical field Hc1 ∼ 1015 G (Baym et al. 1969a). Proton
Andersson & Comer 2006 and references therein for superconductivity is destroyed at the higher critical field
a review of other approaches using a 3+1 spacetime Hc2 ∼ 1016 G (Baym et al. 1969a). Due to entrainment
decomposition). This fully covariant approach not only effects, neutron superfluid vortices carry a magnetic flux
facilitates the comparison with the relativistic theory as well, given by (Sedrakyan & Shakhabasyan 1980;
(Carter et al. 2006; Chamel 2008), but more importantly Alpar et al. 1984)
43 Page 8 of 14 J. Astrophys. Astr. (September 2017) 38:43
 
mp 2012). These results are at variance with those obtained
= 0 −1 , (15) from hydrodynamical studies (Epstein 1988; Sedrakian
mp
1996; Magierski & Bulgac 2004a, b; Magierski 2004;
where 0 = hc/(2e). Electrons scattering off the Martin & Urban 2016). However, as discussed in Martin
magnetic field of the vortex lines leads to a strong & Urban (2016), these approaches are only valid if the
frictional coupling between the core neutron super- neutron superfluid coherence length is much smaller
fluid and the electrically charged particles (Alpar et al. than the typical size of the spatial inhomogeneities, a
1984). Neutron superfluid vortices could also interact condition that is usually not fulfilled in most region
with proton fluxoids (Sauls 1989; Muslimov & Tsy- of the inner crust. The neglect of neutron pairing in
gan 1985; Mendell 1991; Chau et al. 1992), and this the quantum calculations of Chamel (2012) has been
may have important implications for the evolution of recently questioned (Gezerlis et al. 2014; Martin &
the star (Srinivasan 1997; Sauls 1989; Srinivasan et al. Urban 2016). Although detailed numerical calculations
1990; Ruderman 1995; Ruderman et al. 1998; Bhat- are still lacking, the analytical study of Carter et al.
tacharya 2002). For typical neutron star parameters (2005) suggested that neutron pairing is unlikely to have
(P = 10 ms, B = 1012 G, radius R = 10 km), the a large impact on the entrainment coupling.
numbers of neutron superfluid vortices and proton flux-
oids are of the order n υ π R 2 ∼ 1018 and n π R 2 ∼
1030 , respectively. Such large numbers justify a smooth- 4. Observational manifestations
averaged hydrodynamical description of neutron stars.
However, this averaging still requires the understanding 4.1 Pulsar frequency glitches
of the underlying vortex dynamics (Graber et al. 2017).
A more elaborate treatment accounting for the macro- Pulsars are neutron stars spinning very rapidly with
scopic anisotropy induced by the underlying presence extremely stable periods P ranging from milliseconds
of vortices and/or flux tubes was developed by Carter to seconds, with delays Ṗ ≡ dP/dt that in some cases
based on a Kalb-Ramond type formulation (Carter do not do not exceed 10−21 , as compared to 10−18 for
2000) (see also Gusakov & Dommes (2016) and ref- the most accurate atomic clocks (Hinkley et al. 2013).
erences therein). In recent years, simulations of large Nevertheless, irregularities have been detected in long-
collections (∼102 –104 ) of vortices have been carried term pulsar timing observations (Lyne et al. 1995). In
out, thus providing some insight on collective behav- particular, some pulsars have been found to suddenly
iors, such as vortex avalanches (Warszawski & Melatos spin up. These ‘glitches’ in their rotational frequency
2013). However, these simulations have been restricted , ranging from / ∼ 10−9 to ∼10−5 , are gener-
so far to Bose condensates. The extent to which the ally followed by a long relaxation lasting from days
results can be extrapolated to neutron stars remains to be to years, and sometimes accompanied by an abrupt
determined. Such large-scale simulations also require change of the spin-down rate from | / ˙ |
˙ ∼ 10−6 up
microscopic parameters determined by the local dynam- −2
to ∼10 . At the time of this writing, 482 glitches
ics of individual vortices (Bulgac et al. 2013). have been detected in 168 pulsars (Espinoza et al.
The variational formulation of multifluid hydrody- 2011). Since these phenomena have not been observed
namics was extended for studying the magnetoelasto- in any other celestial bodies, they must reflect specific
hydrodynamics of neutron star crusts, allowing for a properties of neutron stars (for a recent review, see,
consistent treatment of the elasticity of the crust, super- Haskell & Melatos (2015)). In particular, giant pul-
fluidity and the presence of a strong magnetic field, both sar frequency glitches / ∼ 10−6 –10−5 as detected
within the Newtonian theory (Carter et al. 2006; Carter in the emblematic Vela pulsar are usually attributed to
& Chachoua 2006) and in the fully relativistic con- sudden transfers of angular momentum from a more
text (Carter & Samuelsson 2006). In particular, these rapidly rotating superfluid component to the rest of the
formulations can account for the entrainment of the neu- star whose rotation frequency is directly observed (for
tron superfluid by the crustal lattice (Carter et al. 2006), a short historical review of theoretical developments,
a non-dissipative effect arising from Bragg scattering of see Chamel (2015) and references therein). The role
unbound neutrons first studied in Carter et al. (2005), of superfluidity is corroborated by the very long relax-
Chamel (2005, 2006) using the band theory of solids. ation times (Baym et al. 1969a) and by experiments
More recent systematic calculations based on a more with superfluid helium (Tsakadze & Tsakadze 1980).
realistic description of the crust have confirmed that The standard scenario of giant pulsar glitches is the fol-
these entrainment effects can be very strong (Chamel lowing. The inner crust of a neutron star is permeated
J. Astrophys. Astr. (September 2017) 38:43 Page 9 of 14 43

by a neutron superfluid that is weakly coupled to the also Page & Reddy (2012)). The thermal relaxation
electrically charged particles by mutual friction forces has been already monitored in a few systems (see
(in a seminal work, Alpar et al. (1984) argued that Waterhouse et al. (2016) and references therein). The
the core neutron superfluid is strongly coupled to the thermal relaxation time depends on the properties of
core, and therefore does not participate to the glitch). the crust, especially the heat capacity. In turn, the onset
The superfluid thus follows the spin-down of the star of neutron superfluidity leads to a strong reduction
via the motion of vortices away from the rotation axis of the heat capacity at temperatures T  Tc thus
unless vortices are pinned to the crust (Anderson & delaying the thermal relaxation of the crust (Fortin
Itoh 1975). In such a case, a lag between the super- et al. 2010). If neutrons were not superfluid, they
fluid and the rest of the star will build up, inducing a could store so much heat that the thermal relaxation
Magnus force acting on the vortices. At some point, the would last longer than what is observed (Shternin
vortices will suddenly unpin, the superfluid will spin et al. 2007; Brown & Cumming 2009). On the other
down and, by the conservation of angular momentum hand, the thermal relaxation of these systems is not
the crust will spin up. During subsequent relaxation, completely understood. For instance, additional heat
vortices progressively repin until the next glitch (Pines sources of unknown origin are needed in order to
& Alpar 1985). This scenario is supported by the anal- reproduce the observations (Waterhouse et al. 2016;
ysis of the glitch data, suggesting that the superfluid Brown & Cumming 2009; Degenaar et al. 2013, 2014;
represents only a few per cent of the angular momen- Turlione et al. 2015; Degenaar et al. 2015; Merritt
tum reservoir of the star (Alpar et al. 1993; Datta & et al. 2016). These discrepancies may also originate
Alpar 1993; Link et al. 1999). On the other hand, this from a lack of understanding of superfluid proper-
interpretation has been recently challenged by the 2007 ties (Turlione et al. 2015). In particular, the low-energy
glitch detected in PSR J1119−6127, and by the 2010 collective excitations of the neutron superfluid were
glitch in PSR B2334+61 (Yuan et al. 2010; Alpar 2011; found to be strongly mixed with the vibrations of
Akbal et al. 2015). More importantly, it has also been the crystal lattice, and this can change substantially
shown that the neutron superfluid in the crust of a neu- the thermal properties of the crust (Chamel et al.
tron star does not contain enough angular momentum 2013, 2016).
to explain giant glitches due to the previously ignored
effects of Bragg scattering (Chamel & Carter 2006; 4.3 Rapid cooling of Cassiopeia A
Andersson et al. 2012; Chamel 2013; Delsate et al.
2016). This suggests that the core superfluid plays a Cassiopeia A is the remnant of a star that exploded
more important role than previously thought (Ho et al. 330 years ago at a distance of about 11000 light
2015; Pizzochero et al. 2017). In particular, the core years from us. It owes its name to its location in
superfluid could be decoupled from the rest of the star the constellation Cassiopeia. The neutron star is not
due to the pinning of neutron vortices to proton flux- only the youngest known, thermally emitting, isolated
oids (Ruderman et al. 1998; Gügercinoğlu & Alpar neutron star in our Galaxy, but it is also the first iso-
2014). So far, most global numerical simulations of pul- lated neutron star for which cooling has been directly
sar glitches have been performed within the Newtonian observed. Ten-year monitoring of this object seems to
theory (Larson & Link 2002; Peralta et al. 2006; Sidery indicate that its temperature has decreased by a few
et al. 2010; Haskell et al. 2012). However, a recent study per cent since its discovery in 1999 (Heinke & Ho
shows that general relativity could significantly affect 2010) (see also the analysis of Elshamouty et al. (2013),
the dynamical evolution of neutron stars (Sourie et al. Posselt et al. (2013) suggesting that the temperature
2017). decline is not statistically significant). If confirmed,
this cooling rate would be substantially faster than
4.2 Thermal relaxation of transiently accreting that expected from nonsuperfluid neutron-star cooling
neutron stars during quiescence theories. It is thought that the onset of neutron super-
fluidity opens a new channel for neutrino emission
In a low-mass X-ray binary, a neutron star accretes from the continuous breaking and formation of neu-
matter from a companion star during several years tron pairs. This process, which is most effective for
or decades, driving the neutron-star crust out of its temperatures slightly below the critical temperature of
thermal equilibrium with the core. After the accre- the superfluid transition, enhances the cooling of the
tion stops, the heated crust relaxes towards equilibrium star during several decades. As a consequence, obser-
(see section 12.7 of Chamel & Haensel (2008), see vations of Cassiopeia A put stringent constraints on the
43 Page 10 of 14 J. Astrophys. Astr. (September 2017) 38:43

critical temperatures of the neutron superfluid and pro- frequencies ranging from 18 Hz to 1800 Hz (see Tur-
ton superconductor in neutron-star cores (Page et al. olla et al. (2015) for a recent review). As anticipated by
2011; Shternin et al. 2011; Ho et al. 2015). However, Duncan (1998), these QPOs are thought to be the sig-
this interpretation has been questioned and alternative natures of global magneto-elastic seismic vibrations of
scenarios have been proposed (Blaschke et al. 2013; the star. If this interpretation is confirmed, the analysis
Negreiros et al. 2013; Sedrakian 2013; Noda et al. 2013; of these QPOs could thus provide valuable information
Bonanno et al. 2014; Ouyed et al. 2015; Sedrakian on the interior of a neutron star. In particular, the iden-
2016; Taranto et al. 2016), most of which still requir- tification of the modes could potentially shed light on
ing superfluidity and/or superconductivity in neutron the existence of superfluid and superconducting phases
stars. (Gabler et al. 2013).

4.4 Pulsar timing noise and rotational evolution

Apart from pulsar frequency glitches, superfluidity and 5. Conclusion


superconductivity may leave their imprint on other
timing irregularities. In particular, pulsar timing noise The existence of superfluid and superconducting phases
(Lyne et al. 1995) could be the manifestation of super- in the dense matter constituting the interior of neutron
fluid turbulence although other mechanisms are likely stars has been corroborated both by theoretical develop-
to play a role (see Melatos & Link (2014) and ref- ments and by astrophysical observations. In particular,
erences therein). Interpreting the long-period (∼100– neutron stars are expected to contain a 1 S0 neutron
1000 days) oscillations in the timing residuals of some superfluid permeating the inner region of the crust and
pulsars such as PSR B1828−11 (Kerr et al. 2016) as the outer core, a 3 PF2 neutron superfluid in the outer
evidence of free precession, it has been argued that core, and a 1 S0 proton superconductor in the outer
either the neutron superfluid does not coexist with the core. Still, many aspects of these phenomena need to
proton superconductor in the core of a neutron star, be better understood. Due to the highly nonlinear char-
or the proton superconductor is type I so as to avoid acter of the pairing mechanism giving rise to nuclear
pinning of neutron superfluid vortices to proton flux- superfluidity and superconductivity, the associated crit-
oids (Link 2003, 2007). However, this conclusion seems ical temperatures remain very uncertain, especially for
premature in view of the complexity of the neutron- the 3 PF2 channel. The dynamics of these phase transi-
star dynamics (Alpar 2005; Glampedakis et al. 2009). tions as the star cools down, and the possible formation
Alternatively, these oscillations might be related to the of topological defects need to be explored. Although
propagation of Tkachenko waves in the vortex lattice the formalism for describing the relativistic smooth-
(see Haskell 2011 and references therein). The presence averaged magnetoelastohydrodynamics of superfluid
of superfluids and superconductors in the interior of a and superconducting systems already exists, modelling
neutron star may also be revealed from the long-term the global evolution of neutron stars in full general rel-
rotational evolution of pulsars by measuring the brak- ativity still remains very challenging. To a large extent,
ing index n = / ¨ ˙ 2 . Deviations from the canonical the difficulty lies in the many different scales involved,
value n = 3 as predicted by a rotating magnetic dipole from the kilometre size of the star down to the size of
model in vacuum can be explained by the decoupling individual neutron vortices and proton fluxoids at the
of the neutron superfluid in the core of a neutron star scale of tens or hundred fermis.
(due to pinning to proton fluxoids for instance) (Alpar & Studies of neutron-star dynamics using the Newto-
Baykal 2006; Ho & Andersson 2012). However, a sim- nian theory provide valuable qualitative insight, and
ilar rotational evolution could be mimicked by other should thus be pursued.
mechanisms without invoking superfluidity (see Pétri The presence of other particles such as hyperons or
(2016) for a recent review). deconfined quarks in the inner core of neutron stars adds
to the complexity. The occurrence of exotic superfluid
4.5 Quasi-periodic oscillations in soft gamma-ray and superconducting phases remains highly speculative
repeaters due to the lack of knowledge of dense matter. On the
other hand, astrophysical observations offer a unique
Quasi-periodic oscillations (QPOs) in the hard X-ray opportunity to probe the phase diagram of matter under
emission were detected in the tails of giant flares from extreme conditions that are inaccessible in terrestrial
SGR 1806-20, SGR 1900+14 and SGR 0526-66, with laboratories.
J. Astrophys. Astr. (September 2017) 38:43 Page 11 of 14 43

Acknowledgements Baym, G., Pethick, C. J., Pines, D., 1969b, Nature, 224, 674.
Bekarevich, I., Khalatnikov, I. 1961, Sov. Phys. JETP, 13,
This work has been supported by the Fonds de la 643.
Recherche Scientifique – FNRS (Belgium) under grant Bhattacharya, D. 2002, J. Astrophys. Astr., 23, 67.
n◦ CDR J.0187.16, and by the European Cooperation in Blaschke, D., Grigorian, H., Voskresensky, D. N. 2013, Phys.
Science and Technology (COST) action MP1304 New- Rev. C, 88, 065805.
Bogoliubov, N. N. 1958, Dokl. Ak. Nauk SSSR, 119, 52.
CompStar.
Bonanno, A., Baldo, M., Burgio, G. F., Urpin, V. 2014, Astron.
Astrophys., 561, L5.
Bose, S. N. 1924, Z. Phys., 26, 178.
References Brown, E. F., Cumming, A. 2009, Astrophys. J., 698, 1020.
Broglia, R. A., Zelevinsky, V. (eds) 2013, Fifty Years of
Abo-Shaeer, J. R., Raman, C., Vogels, J. M., Ketterle, W. Nuclear BCS: Pairing in Finite Systems, World Scientific
2001, Science, 292, 476. Publishing Co. Pte. Ltd.
Abrikosov, A. A. 1957, Sov. Phys. JETP, 5, 1174. Bulgac, A., Forbes, M. M., Sharma, R. 2013, Phys. Rev. Lett.,
Akbal O., Gügercinoğlu, E., Şaşmaz Muş, S., Alpar, M. A. 110, 241102.
2015, Mon. Not. R. Astron. Soc., 449, 933. Cao, L. G., Lombardo, U., Schuck, P. 2006, Phys. Rev. C, 74,
Alford, M. G., Schmitt, A., Rajagopal, K., Schäfer, T. 2008, 064301.
Rev. Mod. Phys., 80, 1455. Carter, B. 1989, in: Relativistic fluid dynamics, Springer-
Alpar, M. A. 2005, in: The Electromagnetic Spectrum of Neu- Verlag, p. 1.
tron Stars, edited by A. Baykal, S. K. Yerli, S. C. Inam & S. Carter, B. 2000, in: Topological Defects and the Non-
Grebenev, NATO Science Series II, Mathematics, Physics equilibrium Dynamics of Symmetry Breaking Phase Tran-
and Chemistry, 210, Springer, Dordrecht, p. 33. sitions, edited by Y. M. Bunkov & H. Godfrin, Springer,
Alpar, M. A. 2011, AIP Conf. Proc., 1379, 166. New York, p. 267.
Alpar, M. A., Baykal, A. 2006, Mon. Not. Roy. Astron. Soc., Carter, B. 2001, in: Lecture Notes in Physics, 578, edited by
372, 489. D. Blaschke, N. K. Glendenning & A. Sedrakian, Springer,
Alpar, M. A., Langer, S. A., Sauls, J. A. 1984, Astrophys. J., p. 54.
282, 533. Carter, B. 1975, Ann. Phys., 95, 53.
Alpar, M. A., Chau, H. F., Cheng, K. S., Pines, D. 1993, Carter, B., Chachoua, E. 2006, Int. J. Mod. Phys. D, 15, 1329.
Astrophys. J., 409, 345. Carter, B., Chamel, N. 2004, Int. J. Mod. Phys. D, 13, 291.
An, Hongjun et al. 2014, Astrophys. J., 790, 60. Carter, B., Chamel, N. 2005a, Int. J. Mod. Phys. D, 14, 717.
Anderson, P. W., Itoh, N. 1975, Nature, 256, 25. Carter, B., Chamel, N. 2005b, Int. J. Mod. Phys. D, 14, 749.
Anderson, M. H., Ensher, J. R., Matthews, M. R., Wieman, Carter, B., Khalatnikov, I. M. 1994, Rev. Math. Phys., 6, 277.
C. E., Cornell, E. A. 1995, Science, 269, 198. Carter, B., Samuelsson, L. 2006, Classical Quant. Grav., 23,
Andersson, N., Comer, G. L. 2006, Classical Quant. Grav. 5367.
23, 5505. Carter, B., Chamel, N., Haensel, P. 2005, Nucl. Phys. A, 748,
Andersson, N., Comer, G. L. 2007, Living Rev. Relativity, 10, 675.
1. Carter, B., Chamel, N., Haensel, P. 2005, Nucl. Phys. A, 759,
Andersson, N., Glampedakis, K., Ho, W. C. G., Espinoza, C. 441.
M. 2012, Phys. Rev. Lett., 109, 241103. Carter, B., Chachoua, E., Chamel, N. 2006, Gen. Rel. Grav.,
Andreev, A. F., Bashkin, E. P. 1976, Sov. Phys. JETP, 42, 38, 83.
164. Carter, B., Chamel, N., Haensel, P. 2006, Int. J. Mod. Phys.
Baldo, M., Schulze, H.-J. 2007, Phys. Rev. C, 75, 025802. D, 15, 777.
Baldo, M., Elgarøy, Ø., Engvik, L., Hjorth-Jensen, M., Chamel, N., Goriely, S., Pearson, J. M. 2013, in: Fifty Years
Schulze, H.-J. 1998, Phys. Rev. C, 58, 1921. of Nuclear BCS: Pairing in Finite Systems, edited by R. A.
Balibar, S. 2014, in: History of Artificial Cold, Scientific, Broglia & V. Zelevinsky, World Scientific Publishing Co.
Technological and Cultural Issues, Boston Studies in the Pte. Ltd., pp. 284–296.
Philosophy and History of Science, 299, Springer, p. 93. Chamel, N., Haensel, P. 2008, Living Rev. Relativity, 11, 10;
Balibar, S. 2007, J. Low Temp. Phys., 146, 441. http://www.livingreviews.org/lrr-2008-10.
Bardeen, J., Cooper, L. N., Schrieffer, J. R. 1957, Phys. Rev., Chamel, N. 2015, in: Proceedings of The Modern Physics of
108, 1175. Compact Stars 2015, PoS(MPCS2015)013.
Basilico, D., Arteaga, D. P., Roca-Maza, X.; Colò, G. 2015, Chamel, N. 2005, Nucl. Phys. A, 747, 109.
Phys. Rev. C, 92, 035802. Chamel, N. 2006, Nucl. Phys. A, 773, 263.
Baym, G., Pethick, C. J., Pines, D., 1969a, Nature, 224, 673. Chamel, N. 2008, Mon. Not. R. Astron. Soc., 388, 737.
Baym, G., Pethick, C. J., Pines, D., Ruderman, M. 1969, Chamel, N. 2012, Phys. Rev. C, 85, 035801.
Nature, 224, 872. Chamel, N. 2013, Phys. Rev. Lett., 110, 011101.
43 Page 12 of 14 J. Astrophys. Astr. (September 2017) 38:43

Chamel, N., Carter, B. 2006, Mon. Not. R. Astron. Soc., 368, Dong, J. M., Lombardo, U., Zuo W. 2013, Phys. Rev. C, 87,
796. 062801(R).
Chamel, N., Haensel, P. 2006, Phys. Rev. C, 73, 045802. Drozdov, A. P., Eremets, M. I., Troyan, I. A., Ksenofontov,
Chamel, N., Naimi, S., Khan, E., Margueron, J. 2007, Phys. V., Shylin, S. I. 2015, Nature, 525, 73.
Rev. C, 75, 055806. Duncan, R. C. 1998, Astrophys. J., 498, L45.
Chamel, N., Goriely, S., Pearson, J. M., Onsi, M. 2010, Phys. Einstein, A. 1925, Sitzungsberichte der Preussischen
Rev. C, 81, 045804. Akademie der Wissenschaften, 1, 3.
Chamel, N., Page, D., Reddy, S. 2013, Phys. Rev. C, 87, Elshamouty, K. G., Heinke, C. O., Sivakoff, G. R., Ho, W. C.
035803. G., Shternin, P. S., Yakovlev, D. G., Patnaude, D. J., David,
Chamel, N., Stoyanov, Zh. K., Mihailov, L. M., Mutafchieva, L. 2013, Astrophys. J., 777, 22.
Y. D., Pavlov, R. L., Velchev, Ch. J. 2015, Phys .Rev. C, Epstein, R. I. 1988, Astrophys. J., 333, 880.
91, 065801. Espinoza, C. M., Lyne, A. G., Stappers, B. W., Kramer, M.
Chamel, N., Pearson, J. M., Fantina, A. F., Ducoin, C., 2011, Mon. Not. Roy. Astron. Soc., 414, 1679; http://www.
Goriely, S., Pastore, A. 2015, Acta Phys. Pol. B, 46, 349. jb.man.ac.uk/pulsar/glitches.html
Chamel, N., Fantina, A. F., Zdunik, J.L., Haensel, P. 2015, Fantina, A. F., Chamel, N., Mutafchieva, Y. D., Stoyanov, Zh.
Phys. Rev. C, 91, 055803. K., Mihailov, L. M., Pavlov, R. L. 2016, Phys .Rev. C, 93,
Chamel, N., Mutafchieva, Y. D., Stoyanov, Zh K., Mihailov, 015801.
L. M., Pavlov, R. L. 2016, J. Phys.: Conf. Ser., 724, Feynman, R. P. 1955, Prog. Low Temp. Phys., 1, 17.
012034. File, J., Mills, R. G. 1963, Phys. Rev. Lett., 10, 93.
Chamel, N., Page, D., Reddy, S. 2016, J. Phys.: Conf. Ser., Fortin, M., Grill, F., Margueron, J., Page, D., Sandulescu, N.
665, 012065. 2010, Phys. Rev. C, 82, 065804.
Chamel, N., Fantina, A. F., Pearson, J. M., Goriely, S. 2017, Gabler, M., Cerdá-Durán, P., Stergioulas, N., Font, J. A.,
EPJ Web of Conferences, 137, 09001. Müller, E. 2013, Phys. Rev. Lett., 111, 211102.
Charbonneau, J., Zhitnitsky, A. 2007, Phys. Rev. C, 76, Gezerlis, A., Pethick, C. J., Schwenk, A. 2014, in: Novel
015801. Superfluids, Volume 2, edited by K. H. Bennemann & J. B.
Chatterjee, D., Vidaña, I. 2016, Eur. Phys. J. A, 52, 29. Ketterson, Oxford University Press, Chapter 22.
Chatterjee, D., Elghozi, T., Novak, J., Oertel, M. 2015, Mon. Ginzburg, V. L., Kirzhnits, D. A. 1964, Zh. Eksp. Teor. Fiz.,
Rot. R. Astron. Soc., 447, 3785. 47, 2006; Sov. Phys. JETP. 20, 1346 (1965).
Chau, H. F., Cheng, K. S., Ding, K. Y. I 1992, Astrophys. J., Ginzburg, V. L. 1969, J. Stat. Phys., 1, 3.
399, 213. Glampedakis, K., Andersson, N., Jones, D. I. 2009, Mon. Not.
Datta, B., Alpar, M. A. 1993, Astron. Astrophys., 275, 210. R. Astron. Soc., 394 ,1908.
Davis, K., Mewes, M.-O., Andrews, M. R., van Druten, N. J., Gorter, C. J. 1955, Prog. Low Temp. Phys., 1, 1.
Durfee, D. S., Kurn, D. M., Ketterle, W. 1995, Phys. Rev. Gourgoulhon, E. 2006, EAS Publications Series, 21, 43.
Lett., 75, 3969. Graber, V., Andersson, N., Hogg, M. 2017, Int. J. Mod. Phys.
Dean, D. J., Hjorth-Jensen, M. 2003, Rev. Mod. Phys., 75, D, 26, 1730015.
607. Gügercinoğlu, E., Alpar, M. A. 2014, Astrophys. J., 788, L11.
Deaver, B., Fairbank, W. 1961, Phys. Rev. Lett., 7, 43. Gusakov, M. E., Dommes, V. A. 2016, Phys. Rev. D, 94,
Degenaar, N., Wijnands, R., Brown, E. F., Altamirano, D., 083006.
Cackett, E. M., Fridriksson, J., Homan, J., Heinke, C. O., Gusakov, M. E., Haensel, P. 2005, Nucl. Phys. A., 761, 333.
Miller, J. M., Pooley, D., Sivakoff, G. R. 2013, Astrophys. Haensel, P., Potekhin, A. Y., Yakovlev, D. G. 2007, Neutron
J., 775, 48. Stars 1: Equation of State and Structure, Springer, New
Degenaar, N., Medin, Z., Cumming, A., Wijnands, R., Wolff, York.
M. T., Cackett, E. M., Miller, J. M., Jonker, P. G., Homan, Hall, H. E. 1960, Adv. Phys., 9, 89.
J., Brown, E. F. 2014, Astrophys. J., 791, 47. Hall, H. E., Vinen, W. F. 1956, Proc. R. Soc. A., 238, 215.
Degenaar, N., Wijnands, R., Bahramian, A., Sivakoff, G. R., Harding, A. K. 2013, Frontiers of Physics, 8, 679.
Heinke, C. O., Brown, E. F., Fridriksson, J. K., Homan, Haskell, B. 2011, Phys. Rev. D, 83, 043006.
J., Cackett, E. M., Cumming, A., Miller, J. M., Altami- Haskell, B., Melatos, A. 2015, Int. J. Mod. Phys. D, 24,
rano, D., Pooley, D. 2015, Mon. Not. R. Astron. Soc., 451, 1530008.
2071. Haskell, B., Pizzochero, P. M., Sidery, T. 2012, Mon. Not. R.
Delsate, T., Chamel, N., Gürlebeck, N., Fantina, A. F., Pear- Astron. Soc., 420, 658.
son, J. M., Ducoin, C. 2016, Phys. Rev. D, 94, 023008. Heinke, C. O., Ho, W. C. G. 2010, Astrophys. J., 719, L167.
Deshpande, A. A., Ramachandran, R., Srinivasan, G. 1995, Hess, G. B., Fairbank, W. M. 1967, Phys. Rev. Lett., 19,
J. Astrophys. Astr., 16, 69. 216.
Ding, D., Rios, A., Dussan, H., Dickhoff, W. H., Witte, S. J., Hinkley, N., Sherman, J. A., Phillips, N. B., Schioppo, M.,
Carbone, A., Polls, A. 2016, Phys. Rev. C, 94, 025802. Lemke, N. D., Beloy, K., Pizzocaro, M., Oates, C. W.,
Doll, R., Näbauer, M. 1961, Phys. Rev. Lett., 7, 51. Ludlow, A. D. 2013, Science, 341, 1215.
J. Astrophys. Astr. (September 2017) 38:43 Page 13 of 14 43

Ho, W. C. G., Andersson, N. 2012, Nature Phys., 8, 787. Meissner, W., Ochsenfeld, R. 1933, Naturwissenschaften, 21,
Ho, W. C. G., Espinoza, C. M., Antonopoulou, D., Andersson, 787.
N. 2015, Science Adv., 1, e1500578. Melatos, A., Link, B. 2014, Mon. Not. R. Astron. Soc., 437,
Ho, W. C. G., Elshamouty, K. G., Heinke, C. O., Potekhin, 21.
A. Y. 2015, Phys. Rev. C, 91, 015806. Mendell, G. 1991, Astrophys. J., 380, 515.
Hoffberg, M., Glassgold, A. E., Richardson, R. W., Ruder- Merritt, R. L., Cackett, E. M., Brown, E. F., Page, D., Cum-
man, M. 1970, Phys. Rev. Lett., 24, 775. ming, A., Degenaar, N., Deibel, A., Homan, J., Miller, J.
Horowitz, C. J., Berry, D. K., Briggs, C. M., Caplan, M. E., M., Wijnands, R. 2016, Astrophys. J., 833, 186.
Cumming, A., Schneider, A. S. 2015, Phys. Rev. Lett., 114, Migdal, A. B. 1959, Nucl. Phys., 13, 655.
031102. Muslimov, A. G., Tsygan, A. I. 1985, Astrophys. Space Sci.,
Kaminker, A. D., Yakovlev, D. G., Potekhin, A. Y., Shibazaki, 115, 43.
N., Shternin, P. S., Gnedin, O. Y. 2007, Astrophys. Space Negreiros, R., Schramm, S., Weber, F. 2013, Phys. Lett. B,
Sci., 308, 423. 718, 1176.
Kapitsa, P. 1938, Nature, 141, 74. Ng, C. Y., Kaspi, V. M. 2011, AIP Conf. Proc., 1379, 60.
Keesom, W. H., Kok, J. A. 1932, Proc. Roy. Acad. Amsterdam, Noda, T., Hashimoto, M., Yasutake, N., Maruyama, T., Tat-
35, 743. sumi, T., Fujimoto, M. 2013, Astrophys. J., 765, 1.
Kerr, M., Hobbs, G., Johnston, S., Shannon, R. M. 2016, Mon. Olausen, S. A., Kaspi, V. M. 2014, Astrophys. J. Suppl. Ser.,
Not. R. Astron. Soc., 455, 1845. 212, 6.
Kheto, A., Bandyopadhyay, D. 2014, Phys. Rev. D, 89, Onsager, L. 1949, Nuovo Cimento, 6, 249.
023007. Osheroff, D. D., Richardson, R. C., Lee, D. M. 1972, Phys.
Kirzhnits, D. A. 1970, Radiophys. Quantum Electron., 13, Rev. Lett., 28, 885.
1424. Ouyed, R., Leahy, D., Koning, N. 2015, Res. Astron. Astro-
Kreim, S., Hempel, M., Lunney, D., Schaffner-Bielich, J. phys., 15, 483.
2013, Int. J. Mass Spec., 349-350, 63. Page, D., Lattimer, J. M., Prakash, M. 2014, in: Novel Super-
Landau, L. 1941, Phys. Rev., 60, 356. fluids: volume 2, edited by K. H. Bennemann & J. B.
Larson, M. B., Link, B. 2002, Mon. Not. R. Astron. Soc., 333, Ketterson, Oxford University Press, p. 505.
613. Page, D., Reddy, S. 2012, in: Neutron Star Crust, edited by C.
Leggett, A. J. 2006, Quantum Liquids: Bose Condensation A. Bertulani & J. Piekarewicz, Nova Science Publishers,
and Cooper Pairing in Condensed-Matter Systems, Oxford New York, p. 281.
University Press. Page, D., Prakash, M., Lattimer, J. M., Steiner, A. W. 2011,
Link, B. 2003, Phys. Rev. Lett., 91, 101101. Phys. Rev. Lett., 106, 081101.
Link, B. 2007, Astrophys. Space Sci., 308, 435. Pastore, A. 2015, Phys. Rev. C, 91, 015809.
Link, B., Epstein, R. I., Lattimer, J. M. 1999, Phys. Rev. Lett., Pastore, A., Margueron, J., Schuck, P., Viñas, X. 2013, Phys.
83, 3362. Rev. C, 88, 034314.
Lombardo, U., Schulze, H.-J. 2001 in: Physics of Neutron Pearson, J. M., Goriely, S., Chamel, N. 2011, Phys. Rev. C,
Star Interiors, edited by D. Blaschke, N. K. Glendenning 83, 065810.
& A. Sedrakian, Springer, p. 30. Peralta, C., Melatos, A., Giacobello, M., Ooi, A. 2006, Astro-
London, F. 1938, Nature, 141, 643. phys. J., 651, 1079.
Lyne, A. G., Pritchard, R. S., Shemar, S. L. 1995, J. Astrophys. Pétri, J. 2016, J. Plasma Phys., 82, 635820502.
Astr., 16, 179. Pili, A. G., Bucciantini, N., Del Zanna L. 2014, Mon. Not. R.
Machleidt, R. 2017, Int. J. Mod. Phys. E, 26, 1740018. Astron. Soc., 439, 3541.
Magierski, P. 2004, Int. J. Mod. Phys. E, 13, 371. Pines, D., Alpar, M. A. 1985, Nature, 316, 27.
Magierski, P., Bulgac, A. 2004, Acta Phys. Pol. B, 35, Pizzochero, P. M., Antonelli, M., Haskell, B., Seveso, S.,
1203. 2017, Nature Astronomy, 1, 0134.
Magierski, P., Bulgac, A. 2004, Nucl. Phys. A, 738, 143. Pons, J. A., Viganò, D., Rea, N. 2013, Nature Physics, 9, 431.
Makishima, K., Enoto, T., Hiraga, J. S., Nakano, T., Posselt, B., Pavlov, G. G., Suleimanov, V., Kargaltsev, O.
Nakazawa, K., Sakurai, S., Sasano, M., Murakami, H. 2013, Astrophys. J., 779, 186.
2014, Phys. Rev. Lett., 112, 171102. Potekhin, A. Y., Pons, J. A., Page, D. 2015, Space Sci. Rev.,
Margueron, J., Sandulescu, N. 2012, in: Neutron Star Crust, 191, 239.
edited by C. Bertulani & J. Piekarewicz, Nova Science Prakash, M., Lattimer, J. M., Pons, J. A., Steiner, A. W.,
Publishers, New York, p. 65. Reddy, S. 2001, in: Lecture Notes in Physics, 578, edited by
Margueron, J., Khan, E. 2012, Phys. Rev. C, 86, 065801. D. Blaschke, N. K. Glendenning & A. Sedrakian, Springer,
Martin, N., Urban, M. 2016, Phys. Rev. C, 94, 065801. p. 364.
Maurizio, S., Holt, J. W., Finelli, P. 2014, Phys. Rev. C, 90, Prix. R. 2004, Phys. Rev. D, 69, 043001.
044003. Prix, R. 2005, Phys. Rev. D, 71, 083006.
43 Page 14 of 14 J. Astrophys. Astr. (September 2017) 38:43

Rea, N., Esposito, P., Turolla, R., Israel, G. L., Zane, S., Stella, Tamagaki, R. 1970, Prog. Theor. Phys., 44, 905.
L., Mereghetti, S., Tiengo, A., Götz, D., Göğüş, E., Kou- Taranto, G., Burgio, G. F., Schulze, H.-J. 2016, Mon. Not. R.
veliotou, C. 2010, Science, 330, 944. Astron. Soc., 456, 1451.
Regal, C. A., Greiner, M., Jin, D. S. 2004, Phys. Rev. Lett., Thompson, C., Duncan, R. C. 1992, Astrophys. J., 392,
92, 040403. L9.
Reppy, J. D., Depatie, D. 1964, Phys. Rev. Lett., 12, 187. Tiengo, A. et al. 2013, Nature, 500, 312.
Rjabinin, J. N., Shubnikov, L. W. 1935a, Phys. Zeit. Swo. Tisza, L. 1938, Nature, 141, 913.
Union, 7, 122. Tkachenko, V. K. 1966, Sov. Phys. JETP, 22, 1282.
Rjabinin, J. N., Shubnikov, L. W. 1935b, Nature, 135, 581. Tsakadze, J. S., Tsakadze, S. J. 1980, J. Low Temp. Phys., 39,
Ruderman, M. 1995, J. Astrophys. Astr., 16, 207. 649.
Ruderman, M., Zhu, T., Chen, K. 1998, Astrophys. J., 492, Turlione, A., Aguilera, D. N., Pons, J. A. 2015, Astron. Astro-
267. phys., 577, A5.
Samuelsson L., Andersson, N. 2007, Mon. Not. Roy. Astr. Turolla, R., Zane, S., Watts, A. L. 2015, Rep. Prog. Phys., 78,
Soc., 374, 256. 116901.
Sauls, J. 1989, in: NATO Advanced Science Institutes (ASI) Utama, R., Piekarewicz, J., Prosper, H. B. 2016, Phys. Rev.
Series C, edited by H. Ögelman & E. P. J. van den C, 93, 014311.
Heuvel, NATO Advanced Science Institutes, Kluwer Aca- van Delft, D., Kes, P. 2010, Phys. Today, 63, 9, 38
demic/Plenum Press, New York, vol. 262, p. 457. Vardanyan, G. A., Sedrakyan, D. M. 1981, Sov. Phys. JETP,
Schuck, P. 2014, J. Phys.: Conf. Series, 569, 012012. 54, 919.
Sedrakian, A., Clark, J. W. 2006, in: Pairing in Fermionic Vietri, M., Stella, L., Israel, G. L. 2007, Astrophys. J., 661,
Systems, World Scientific Publishing, p. 135. 1089.
Sedrakian, A. 1996, Astrophys. Space Sci., 236, 267. Warszawski, L., Melatos, A. 2013, Mon. Not. Roy. Astron.
Sedrakian, A. 2010, Acta Phys. Polon. B: Proc. Suppl., 3, Soc., 428, 1911.
669. Watanabe, G., Maruyama, T. 2012, in: Neutron Star Crust,
Sedrakian, A. 2013, Astron. Astrophys., 555, L10. edited by C. Bertulani & J. Piekarewicz, Nova Science
Sedrakian, A. 2016, Eur. Phys. J. A, 52, 44. Publishers, p. 23.
Sedrakyan, D. M., Shakhabasyan, K. M. 1980, Astrofizika, Waterhouse, A. C., Degenaar, N., Wijnands, R., Brown, E.
16, 727. F., Miller, J. M., Altamirano, D., Linares, M. 2016, Mon.
Sharma, B. K., Centelles, M., Viñas, X., Baldo, M., Burgio, Not. R. Astron. Soc., 456, 4001.
G. F. 2015, Astron. Astrophys., 584, A103. Watts, A. L., Strohmayer, T. E. 2007, Adv. Space Res., 40,
Shimizu, K., Kimura, T., Furomoto, S., Takeda, K., Kontani, 1446.
K., Onuki, Y., Amaya, K. 2001, Nature, 412, 316. Wigner, E. P., Seitz, F. 1933, Phys. Rev., 43, 804.
Shternin, P. S., Yakovlev, D. G., Haensel, P., Potekhin, A. Y. Wlazłowski, G., Sekizawa, K., Magierski, P., Bulgac,
2007, Mon. Not. R. Astron. Soc., 382, L43. A., McNeil Forbes, M. 2016, Phys. Rev. Lett., 117,
Shternin, P. S., Yakovlev, D. G., Heinke, C. O., Ho, W. C. G., 232701.
Patnaude, D. J. 2011, Mon. Not. R. Astron. Soc., 412, L108. Wolf, R. A. 1966, Astrophys. J., 145, 834.
Sidery, T., Passamonti, A., Andersson, N. 2010, Mon. Not. R. Wolf, R. N. et al. 2013, Phys. Rev. Lett., 110, 041101.
Astron. Soc., 405, 1061. Woods, P. M., Thompson, C. 2006, in: Compact Stel-
Sourie, A., Oertel, M., Novak, J. 2016, Phys. Rev. D, 93, lar X-ray Sources, edited by W. H. G. Lewin & M.
083004. van der Klis, Cambridge University Press, Cambridge,
Sourie, A., Chamel, N., Novak, J., Oertel, M. 2017, Mon. Not. p. 547.
R. Astron. Soc., 464, 4641. Yarmchuk, E. J., Gordon, M. J. V., Packard, R. E. 1979, Phys.
Srinivasan, G. 1997, in: Stellar Remnants, edited by G. Rev. Lett., 43, 214.
Meynet & D. Schaerer, Springer-Verlag, p. 97. Yu, Y., Bulgac, A. 2003, Phys. Rev. Lett., 90, 161101.
Srinivasan, G., Bhattacharya, D., Muslimov, A. G., Tsygan, Yuan, J. P., Manchester, R. N., Wang, N., Zhou, X.,
A. J. 1990, Current Science, 59, 31. Liu, Z. Y., Gao, Z. F. 2010, Astrophys. J., 719,
Stein, M., Sedrakian, A., Huang, Xu-Guang, Clark, J. W. L111.
2014, Phys. Rev. C, 90, 065804. Zhou, X-R, Schulze, H.-J., Zhao, E-G, Pan, F., Draayer, J. P.
Stein, M., Sedrakian, A., Huang, Xu-Guang, Clark, J. W. 2004, Phys. Rev. C, 70, 048802.
2016, Phys. Rev. C, 93, 015802. Zwierlein, M. W., Abo-Shaeer, J. R., Schirotzek, A.,
Stella, L., DallOsso, S., Israel, G. L. 2005, Astrophys. J., 634, Schunck, C. H., Ketterle, W. 2005, Nature, 435,
L165. 1047.

You might also like